Skip to main content
Log in

Estimation and clustering for partially heterogeneous single index model

  • Regular Article
  • Published:
Statistical Papers Aims and scope Submit manuscript

Abstract

In this paper, our goal is to estimate the homogeneous parameter and cluster the heterogeneous parameters in a partially heterogeneous single index model (PHSIM). To achieve the goal, the minimization criterion for such a single index model is first transformed into a least-squares optimization problem in the population form. Based on the least-squares objective function, we introduce an empirical version for the PHSIM. By minimizing such an empirical version, we estimate the homogeneous parameter and the subgroup-averages of the heterogeneous index directions, and then use a fusion penalized method to identify the subgroup structure of the PHSIM. By the proposed methodologies, the homogeneous parameter and the heterogeneous index directions can be consistently estimated, and the heterogeneous parameters can be consistently clustered. Moreover, the new clustering procedure is simple and robust. Simulation studies are carried out to examine the performance of the proposed methodologies.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Boyd S, Parikh N, Chu E (2011) Distributed optimization and statistical learning via the alternating direction method of multipliers. Foundations Trends Mach Learn 3(1):1–122

    Article  Google Scholar 

  • Chen G, You J (2005) An asymptotic theory for semiparametric generalized least squares estimation in partially linear regression models. Stat Pap 46(2):173–193

    Article  MathSciNet  Google Scholar 

  • Chi EC, Lange K (2015) Splitting methods for convex clustering. J Comput Graph Stat 24(4):994–1013

    Article  MathSciNet  Google Scholar 

  • Cook RD, Ni L (2005) Sufficient dimension reduction via inverse regression: a minimum discrepancy approach. J Am Stat Assoc 100(470):410–428

    Article  MathSciNet  Google Scholar 

  • Engle RF, Granger CW, Rice J, Weiss A (1986) Semiparametric estimates of the relation between weather and electricity sales. J Am Stat Assoc 81(394):310–320

    Article  Google Scholar 

  • Fan J, Li R (2001) Variable selection via nonconcave penalized likelihood and its oracle properties. J Am Stat Assoc 96(456):1348–1360

    Article  MathSciNet  Google Scholar 

  • Hall P, Li KC (1993) On almost linearity of low dimensional projections from high dimensional data. Ann Stat 21(2):867–889

    MathSciNet  MATH  Google Scholar 

  • Kravitz RL, Duan N, Braslow J (2004) Evidence-based medicine, heterogeneity of treatment effects, and the trouble with averages. Milbank Q 82(4):661–687

    Article  Google Scholar 

  • Lagakos SW (2006) The challenge of subgroup analyses-reporting without distorting. N Engl J Med 354(16):1667

    Article  Google Scholar 

  • Lee ER, Noh H, Park BU (2014) Model selection via Bayesian information criterion for quantile regression models. J Am Stat Assoc 109(505):216–229

    Article  MathSciNet  Google Scholar 

  • Lin L, Dong P, Song Y, Zhu L (2017) Upper expectation parametric regression. Stat Sin 27(3):1265–1280

    MathSciNet  MATH  Google Scholar 

  • Ma S, Huang J, Zhang Z, Liu M (2019) Exploration of heterogeneous treatment effects via concave fusion. Int J Biostat 16(1):

  • Radchenko P, Mukherjee G (2017) Convex clustering via l1 fusion penalization. Stat Methodol Ser B 79(5):1527–1546

    Article  MathSciNet  Google Scholar 

  • Rand WM (1971) Objective criteria for the evaluation of clustering methods. J Am Stat Assoc 66(336):846–850

    Article  Google Scholar 

  • Rothwell PM (2005) Subgroup analysis in randomised controlled trials: importance, indications, and interpretation. Lancet 365(9454):176–186

    Article  Google Scholar 

  • Schwarz G (1978) Estimating the dimension of a model. Ann Stat 6(2):461–464

    Article  MathSciNet  Google Scholar 

  • Tibshirani R, Saunders M, Rosset S, Zhu J, Knight K (2005) Sparsity and smoothness via the fused lasso. J R Stat Soc Ser B (Statistical Methodology)

  • Van der Vaart AW (2000) Asymptotic statistics, vol 3. Cambridge University Press, Cambridge

    Google Scholar 

  • Wang H, Li R, Tsai CL (2007) Tuning parameter selectors for the smoothly clipped absolute deviation method. Biometrika 94(3):553–568

    Article  MathSciNet  Google Scholar 

  • Wang T, Xu PR, Zhu LX (2012) Non-convex penalized estimation in high-dimensional models with single-index structure. J Multivar Anal 109:221–235

    Article  MathSciNet  Google Scholar 

  • Wang T, Zhang J, Liang H, Zhu L (2015) Estimation of a groupwise additive multiple-index model and its applications. Stat Sin 25(2):551–566

    MathSciNet  MATH  Google Scholar 

  • Xie H, Huang J (2009) SCAD-penalized regression in high-dimensional partially linear models. Ann Stat 37(2):673–696

    Article  MathSciNet  Google Scholar 

  • Zhang CH (2010) Nearly unbiased variable selection under minimax concave penalty. Ann Stat 38(2):894–942

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lu Lin.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The research was supported by NNSF Projects (11971265, 11901356) of China.

Appendix

Appendix

Proof of Proposition 1

According to the Proposition 2 in Wang et al. (2015), under conditions (B0),(B1)and (B2), there exist constants \(\phi _1\ne 0\) and \(\phi _2\ne 0\) such that \(\varvec{\rho }_{LS}=(\phi _1\varvec{\beta }^T,\phi _2\varvec{\theta }^T)^T\). In the spirit of the justification of Theorem 2 of Wang et al. (2015), we only give a sketch to prove \(\phi _1=1\).

(I) Note that

$$\begin{aligned} \begin{aligned} \varvec{\rho }_{LS}&={\begin{pmatrix}\Sigma _{X,X}&{}\Sigma _{X,Z}\\ \Sigma _{Z,X}&{}\Sigma _{Z,Z}\\ \end{pmatrix}}^{-1} \begin{pmatrix}\text{ Cov }(X,Y)\\ \text{ Cov }(Z,Y) \end{pmatrix}. \end{aligned} \end{aligned}$$
(6.1)

When the condition (A1) of the Proposition 1 holds,

$$\begin{aligned} \begin{aligned} \varvec{\rho }_{LS}&={\begin{pmatrix}\Sigma _{X,X}&{}0\\ 0&{}\Sigma _{Z,Z}\\ \end{pmatrix}}^{-1} \begin{pmatrix}\Sigma _{X,X}^T\varvec{\beta }\\ E[Zg_0(Z^T\varvec{\theta })] \end{pmatrix}\\&=\begin{pmatrix}\varvec{\beta }\\ \Sigma _{Z,Z}^{-1}E[Zg_0(Z^T\varvec{\theta })] \end{pmatrix}. \end{aligned} \end{aligned}$$
(6.2)

Clearly, we have \(\phi _1=1\).

(II) Under the condition (A2) of the Proposition 1, define

$$\begin{aligned} \begin{aligned} S&={\begin{pmatrix}\varvec{\beta }&{}0_{p\times 1}\\ 0_{q\times 1}&{}\varvec{\theta } \end{pmatrix}}, \end{aligned} \end{aligned}$$
(6.3)

and

$$\begin{aligned} \begin{aligned} \phi =(S^T\Sigma S)^{-1}S^T{\begin{pmatrix}E(XY)\\ E(ZY) \end{pmatrix}}, \end{aligned} \end{aligned}$$
(6.4)

where \(\Sigma ={\begin{pmatrix}\Sigma _{X,X}&{}\Sigma _{X,Z}\\ \Sigma _{Z,X}&{}\Sigma _{Z,Z}\\ \end{pmatrix}}\). Because the link function \(g_0(\cdot )\) is analytic, it suffices to consider the case where \(g_0(t)=\Sigma _{k=1}^{\infty }a_kt^k\). Under the condition (A2) of the Proposition 1, the bivariate random vector \((\tilde{X},\tilde{Z})=(X^T\varvec{\beta }, Z^T\varvec{\theta })\) has an elliptically symmetric distribution with mean vector zero. It follows that \(E(\tilde{Z}^{k+1})=0\) if k is even; \(E(\tilde{Z}^{k+1})=c_r\{E(\tilde{Z}^2)\}^r\) and \(E(\tilde{X}\tilde{Z}^k)=c_rE(\tilde{X}\tilde{Z})\{E(\tilde{Z}^2)\}^{r-1}\) for some constant \(c_r\) if \(k+1=2r\). As a result, we have

$$\begin{aligned} \sum _{k+1=2r}a_kc_rE(\tilde{X}\tilde{Z})\{E(\tilde{Z}^2)\}^{r-1}=\phi _2E(\tilde{X}\tilde{Z}), \sum _{k+1=2r}a_kc_r\{E(\tilde{Z}^2)\}^r=\phi _2E(\tilde{Z}^2), \end{aligned}$$
(6.5)

where the summation \(\sum _{k+1=2r}\) is taken over the set \(\{k:k\ge 1, k+1=2r\}.\) Therefore, we can take \(\phi _2=\sum _{k+1=2r}a_kc_r\{E(\tilde{Z}^2)\}^{r-1}.\) We have completed the proof of the Proposition 1. \(\square \)

Proof of Theorem 1

Note that \(\varvec{\theta }\) is heterogeneous. By objective function (2.4), we have the following empirical versions:

$$\begin{aligned} \overline{\mathbf {X} \mathbf {y}}= & {} \overline{\mathbf {X} \mathbf {X}^T}\varvec{\beta }+\frac{1}{n} \sum _{i=1}^n \mathbf {x}_i\mathbf {z}_i^T\varvec{\theta }_i+\frac{1}{n} \sum _{i=1}^n \mathbf {x}_i\varepsilon _i,\\ \overline{\mathbf {Z} \mathbf {y}}= & {} \overline{\mathbf {Z} \mathbf {X}^T}\varvec{\beta }+\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i+\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\varepsilon _i. \end{aligned}$$

Then

$$\begin{aligned}&\overline{\mathbf {X} \mathbf {y}}-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\overline{\mathbf {Z} \mathbf {y}}\\&\quad =\Big (\overline{\mathbf {X} \mathbf {X}^T}-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\overline{\mathbf {Z} \mathbf {X}^T}\Big )\varvec{\beta } +\frac{1}{n} \sum _{i=1}^n \mathbf {x}_i\mathbf {z}_i^T\Big (\varvec{\theta }_i-(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\Big )\\&\qquad +\frac{1}{n} \sum _{i=1}^n\Big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\mathbf {z}_i\Big )\varepsilon _i. \end{aligned}$$

By the result above and estimator (2.7), we get

$$\begin{aligned} \widehat{\varvec{\beta }}-\varvec{\beta }= & {} \Big (\overline{\mathbf {X} \mathbf {X}^T}-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\overline{\mathbf {Z} \mathbf {X}^T}\Big )^{-1}\frac{1}{n} \sum _{i=1}^n \mathbf {x}_i\mathbf {z}_i^T\Big (\varvec{\theta }_i-(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\Big )\\&+\Big (\overline{\mathbf {X} \mathbf {X}^T}-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\overline{\mathbf {Z} \mathbf {X}^T}\Big )^{-1}\frac{1}{n} \sum _{i=1}^n\Big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\mathbf {z}_i\Big )\varepsilon _i. \end{aligned}$$

Because

$$\begin{aligned} \overline{\mathbf {X} \mathbf {X}^T}-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\overline{\mathbf {Z} \mathbf {X}^T}\xrightarrow {P}\Phi =E[XX^T]-E[XZ^T]\big (E[ZZ^T]\big )^{-1}E[ZX^T], \end{aligned}$$

we can obtain that \(\sqrt{n}\Phi (\widehat{\varvec{\beta }}-\varvec{\beta })\) is asymptotically identically distributed as

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T\Omega _i+\frac{1}{\sqrt{n}}\sum _{i=1}^n\Big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\mathbf {z}_i\Big )\varepsilon _i, \end{aligned}$$
(6.6)

where the weight \(\Omega _i=\Big (\varvec{\theta }_i-(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\Big )\). Note that \((\mathbf {x}_i^T, \mathbf {z}_i^T)^T, i=1, \cdots , n\), are independent and identically distributed as \((X^T,Z^T)^T\). We get

$$\begin{aligned}&E\bigg (\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\bigg )=\frac{1}{n} \sum _{i=1}^n E\big (\mathbf {z}_i\mathbf {z}_i^T\big )\varvec{\theta }_i=E\big (ZZ^T\big )\frac{1}{n} \sum _{i=1}^n\varvec{\theta }_i,\\&\quad Cov\bigg (\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\bigg )=\frac{1}{n^2}\sum _{i=1}^nCov\big (\mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i\big )=\frac{1}{n^2}\sum _{i=1}^n\varvec{\theta }_i^T\Delta \varvec{\theta }_i=O(1\sqrt{n}), \end{aligned}$$

where \(\Delta =E\bigg (\big (ZZ^T-E(ZZ^T)\big )\big (ZZ^T-E(ZZ^T)\big )\bigg )\). The above indicates that

$$\begin{aligned} \frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i=E\big (ZZ^T\big )\frac{1}{n} \sum _{i=1}^n\varvec{\theta }_i+O_p(1/\sqrt{n}) \end{aligned}$$

and then \(\Omega _i\xrightarrow {P}(\varvec{\theta }_i-\overline{\varvec{\theta }})\). Thus, the weighted sum \(\frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T\Omega _i\) can be expressed as

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})+o_p(1)\frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }}) \end{aligned}$$

and then is asymptotically identically distributed as \(\frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})\). It can be seen that

$$\begin{aligned} E\bigg (\frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})\bigg )=0 \end{aligned}$$

and

$$\begin{aligned}&Var\bigg (\frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})\bigg )\\&\quad =\frac{1}{n}\sum _{i=1}^n(\varvec{\theta }_i-\overline{\varvec{\theta }})^TE\bigg (\big (ZX^T-E(ZX^T)\big )\big (ZX^T-E(ZX^T)\big )^T\bigg )(\varvec{\theta }_i-\overline{\varvec{\theta }}). \end{aligned}$$

Therefore,

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n\mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})\xrightarrow {D}N(0,\Psi ). \end{aligned}$$
(6.7)

In addition, we can also get that

$$\begin{aligned} \frac{1}{\sqrt{n}} \sum _{i=1}^n\Big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\mathbf {z}_i\Big )\varepsilon _i\xrightarrow {D}N(0,\sigma ^2\Lambda ), \end{aligned}$$
(6.8)

where \(\Lambda =Var\bigg (X-E\big (XZ^T\big )\big (E(ZZ^T)\big )^{-1}Z\bigg )\). By (6.6), (6.7), (6.8) and the uncorrelation between (XZ) and \(\varepsilon \), we get that

$$\begin{aligned} \sqrt{n}(\widehat{\varvec{\beta }}-\varvec{\beta })\xrightarrow {D}N(0, \Phi ^{-1}(\Psi +\sigma ^2\Lambda )\Phi ^{-1}). \end{aligned}$$
(6.9)

Thus, the theorem is proved. \(\square \)

Proof of Theorem 2

We only prove the first result. The proof of the second result is similar. Based on the objective function (2.4), the estimator (2.8) can be denoted as the following empirical form:

$$\begin{aligned} \begin{aligned} \widehat{\overline{\varvec{\theta }}}&=\big (\overline{\mathbf {Z} \mathbf {Z}^T}\big )^{-1}\big (\overline{\mathbf {Z} \mathbf {y} }-\overline{\mathbf {Z} \mathbf {X}^T}\widehat{\varvec{\beta }}\big )\\&=\big (\overline{\mathbf {Z} \mathbf {Z}^T}\big )^{-1}\bigg (\overline{\mathbf {Z} \mathbf {X}^T}\varvec{\beta }+\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T\varvec{\theta }_i+\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\varepsilon _i-\overline{\mathbf {Z} \mathbf {X}^T}\widehat{\varvec{\beta }}\bigg )\\&=\overline{\varvec{\theta }}+\big (\overline{\mathbf {Z} \mathbf {Z}^T}\big )^{-1}\bigg (\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})+\frac{1}{n} \sum _{i=1}^n \mathbf {z}_i\varepsilon _i-\overline{\mathbf {Z} \mathbf {X}^T}(\widehat{\varvec{\beta }}-\varvec{\beta })\bigg ). \end{aligned} \end{aligned}$$
(6.10)

From the proof of the Theorem 1, \(\sqrt{n}\Phi (\widehat{\varvec{\beta }}-\varvec{\beta })\) is asymptotically identically distributed as

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n \mathbf {x}_i\mathbf {z}_i^T(\varvec{\theta }_i-\overline{\varvec{\theta }})+\frac{1}{\sqrt{n}}\sum _{i=1}^n\big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\mathbf {z}_i\big )\varepsilon _i. \end{aligned}$$

By the above results, we can easily see that \(\sqrt{n}(\widehat{\overline{\varvec{\theta }}}-\overline{\varvec{\theta }})\) is asymptotically identically distributed as

$$\begin{aligned}&(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\frac{1}{\sqrt{n}}\sum _{i=1}^n\Big (\mathbf {z}_i\mathbf {z}_i^T-\overline{\mathbf {Z} \mathbf {X}^T}\Phi ^{-1}\mathbf {x}_i\mathbf {z}_i^T\Big )(\varvec{\theta }_i-\overline{\varvec{\theta }})\\&\quad +(\overline{\mathbf {Z} \mathbf {Z}^T})^{-1}\Bigg (\frac{1}{\sqrt{n}}\sum _{i=1}^n\bigg (\mathbf {z}_i-\overline{\mathbf {Z} \mathbf {X}^T}\Phi ^{-1}\Big (\mathbf {x}_i-\overline{\mathbf {X} \mathbf {Z}^T}\big (\overline{\mathbf {Z} \mathbf {Z}^T}\big )^{-1}\mathbf {z}_i\Big )\bigg )\varepsilon _i\Bigg ). \end{aligned}$$

Thus,

$$\begin{aligned} \sqrt{n}(\widehat{\overline{\varvec{\theta }}}-\overline{\varvec{\theta }})\xrightarrow {D}N\Big (0, \big (E(ZZ^T)\big )^{-1}\big (\Gamma +\sigma ^2\Upsilon \big )\big (E(ZZ^T)\big )^{-1}\Big ). \end{aligned}$$
(6.11)

\(\square \)

Proof of Theorem 3

For given \(\beta ^0\), based on the sample \(W_i(\beta ^0)=\mathbf {z}_i(y_i-\mathbf {x}_i^T\beta ^0), i=1,\cdots ,n\) of \(W(\beta ^0)\), the corresponding estimators of \(L, R, G_{L,R}, \cdots \), are denoted by \(\tilde{L}(\beta ^0), \tilde{R}(\beta ^0), \tilde{G}_{L,R}(\beta ^0), \cdots ,\) respectively. Using the sample \(\widehat{W}_i=\mathbf {z}_i(y_i-\mathbf {x}_i\widehat{\beta }), i=1,\cdots ,n\), the corresponding estimators are denoted by \(\widehat{L}, \widehat{R}, \widehat{G}_{L,R}, \cdots \), respectively. By the asymptotic property of the estimator of \(\widehat{\beta }\) given in Theorem 1, and the continuous-mapping theorem for strong consistency (see, e.g., Van der Vaart (2000)), we have \(\widehat{L}-\tilde{L}(\beta ^0)\xrightarrow {a.s.}0, \widehat{R}-\tilde{R}(\beta ^0)\xrightarrow {a.s.}0\) and \(\widehat{G}_{L,R}-\tilde{G}_{L,R}(\beta ^0)\xrightarrow {a.s.}0, \cdots \). Then, for any \(\varepsilon >0\) and for all sufficiently large n, the following inequalities hold almost everywhere:

$$\begin{aligned} |\widehat{L}-\tilde{L}(\beta ^0)|<\varepsilon ,| \widehat{R}-\tilde{R}(\beta ^0)|<\varepsilon ,|\widehat{G}_{L,R}-\tilde{G}_{L,R}(\beta ^0)|<\varepsilon ,\cdots . \end{aligned}$$

This implies that the estimator \(\widehat{L}, \widehat{R}, \widehat{G}_{L,R}, \cdots \), can be replaced by \(\tilde{L}(\beta ^0), \tilde{L}(\beta ^0), \tilde{R}(\beta ^0), \tilde{G}_{L,R}(\beta ^0),\cdots \), respectively, in the proof procedure (see Lemma 1-Lemma 3 in supplementary Material of Radchenko and Mukherjee 2017). Then, based on the population \(W(\beta ^0)\), its expectation \(\alpha \) and the sample from \(W(\beta ^0)\), by the proof of Theorem 1 of Radchenko and Mukherjee (2017), we can prove Theorem 3. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, F., Lin, L., Liu, L. et al. Estimation and clustering for partially heterogeneous single index model. Stat Papers 62, 2529–2556 (2021). https://doi.org/10.1007/s00362-020-01203-2

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00362-020-01203-2

Keywords

Navigation